arXiv:1408.1426v2 [math.PR] 24 Aug 2014

A NOTE ON THE SHARP Lp -COVERGENCE RATE OF UPCROSSINGS TO
THE BROWNIAN LOCAL TIME
ALBERTO OHASHI AND ALEXANDRE B. SIMAS
arXiv:1408.1426v2 [math.PR] 24 Aug 2014
Abstract. In this note, we prove a sharp Lp -rate of convergence of the number of upcrossings
to the local time of the Brownian motion. In particular, it provides novel p-variation estimates
(2 < p < ∞) for the number of upcrossings of the Brownian motion. Our result complements the
fundamental work of Koshnevisan [10] who obtains an almost sure exact rate of convergence in the
sup norm.
1. Introduction
x
Fix x ∈ R, and let (Ω, F, P ) be the Wiener space of the canonical Brownian motion starting from
x, i.e., Ω := C([0, +∞); R) is the space of continuous functions from [0, +∞) to R, F = (Ft )t≥0 is
the natural filtration generated by the Brownian motion satisfying the usual conditions, and Px is the
Wiener measure of the Brownian motion B(ω, t) = ω(t) with Px {B(0) = x} = 1. When x = 0, we just
write P. We endow the Wiener space with the filtration generated by the Brownian motion satisfying
the usual conditions.
The occupation measure of B up to the instant t is the measure µt defined by the relation
µt (A) :=
Z
t
11A (B(s))ds; A ∈ B(R),
0
where B(R) is the Borel sigma algebra of R. In a landmark work, L´evy established that for almost
all trajectories of the Brownian motion and for any t, the random measure µt has a density and
by the classical Trotter Theorem we know that it admits a jointly continuous version {ℓx (t); (x, t) ∈
R × [0, ∞)}, the so-called local-time of the Brownian motion.
Several approximation schemes exist in the literature for the Brownian
R t local time. For instance,
let {ψε (·); ε > 0} be an approximation to the identity. Then limε→0 0 ψε (B(s) − x)ds = ℓx (t) a.s
uniformly over all x ∈ R. See e.g Borodin [4] and Karatzas and Shreve [9] for further references on this
topic. In a different direction, very appealing strong approximation schemes can be constructed from
several types of random walks based on the same probability space. See Bass and Koshnevisan [2]
for further references. The L´evy excursion theory provides other approximations schemes by means
of the number of upcrossings or of the excursions before a given time.
Sharp rates of almost sure convergence of the number of upcrossings to the Brownian local time
in the sup norm and mean squared error are by now well understood. See the fundamental works of
Borodin [4], Koshnevisan [10] and Knight [12]. However, much little is known about Lp (P)-convergence
rates for upcrossings to the Brownian local-time (see e.g Blandine and Vallois [3] for Lp -rates in the
sense of regularization). The goal of this short note is to present sharp rates of convergence for the
number upcrossings to the Brownian local-time in the Lp (P)-sense.
Our main motivation in studying sharp Lp (P)-rates of convergence of the number of upcrossings to
the Brownian local time lies in Itˆ
o formulas for path-dependent functionals of the Brownian motion.
Le˜ao, Ohashi and Simas [14] have recently proved that under suitable p-variation regularity (in the
Date: August 26, 2014.
Key words and phrases. Brownian motion, Local-Time, upcrossings.
We would like to thank professor Davar Koshnevisan for very helpful discussions about this article.
1
2
ALBERTO OHASHI AND ALEXANDRE B. SIMAS
sense of rough path, see e.g [7]) of a non-anticipative Brownian functional Ft : C([0, t]; R) → R, the
process
(1.1)
Ft (Bt ) −
1
2
Z tZ
0
+∞
∂x Fs (t(Bs , x))d(s,x) ℓx (s), 0 ≤ t ≤ T,
−∞
is a Brownian semimartingale, where Bt := {B(s); 0 ≤ s ≤ t} is the Brownian path and 0 < T < ∞ is
a fixed terminal time. Here the d(s,x) ℓ-integral is the pathwise 2D Young integral (see e.g [17, 7, 15])
composed with the Brownian local-time {ℓx (s); 0 ≤ s ≤ T, x ∈ R} and, roughly speaking, a suitable
“space” derivative of F composed with a “terminal value modification” t(Bt , x) of the Brownian paths
(see [14] for further details).
One important step in the proof of (1.1) is a sharp Lp (P)-convergence rate of the number of
upcrossings of the embedded random walk introduced by Knight [11] to the Brownian local time. In
particular, p-variation regularity of the number of upcrossings plays a key role on the existence of the
semimartingale decomposition (1.1) and it is an almost immediate corollary of the main result of this
note.
2. Preliminaries
At first, let us recall the F. Knight [11] construction of an 2−k Z-valued simple symmetric random
walk using a single Brownian motion. For a fixed positive integer k, we define T0k := 0 a.s. and
k
k
)| = 2−k },
< t < ∞; |B(t) − B(Tn−1
Tnk := inf{Tn−1
k
n ≥ 1.
{B(Tnk ); n
k
Then the discrete-time process R :=
≥ 0} is a simple symmetric random walk. For
simplicity of exposition, we are going to imbed R into a continuous-time process. We define Ak as
follows
Ak (t) :=
∞
X
2−k ηnk 11{Tnk ≤t} ; t ≥ 0,
n=1
where
ηnk
k
) = 2−k and Tnk < ∞
1; if B(Tnk ) − B(Tn−1
k
k
−1; if B(Tn ) − B(Tn−1 ) = −2−k and Tnk < ∞
:=

0; if Tnk = ∞,


for k, n ≥ 1. Then, for each k ≥ 1, Ak is a bounded variation martingale w.r.t its natural filtration
Fk . See e.g [13] for details. In the sequel, for a given x ∈ R, let jk (x) be the unique integer such that
(jk (x) − 1)2−k < x ≤ jk (x)2−k . Let us define
o
n
k
u(jk (x)2−k , k, t) := # n ∈ {0, . . . , N k (t) − 1}; Ak (Tnk ) = (jk (x) − 1)2−k , Ak (Tn+1
) = jk (x)2−k ;
for x ∈ R, k ≥ 1, 0 ≤ t < ∞. Here, N k (t) := max{n; Tnk ≤ t} is the length of the embedded random
walk until time t. By the very definition, u(jk (x)2−k , k, t) := number of upcrossings of Ak from (jk (x)−
1)2−k to jk (x)2−k before time t. In the sequel, we are going to denote
U k (t, x) := 22−k u(jk (x)2−k , k, t); (t, x) ∈ [0, +∞) × R.
One fundamental result due to Koshnevisan (see Th. 1.4 and Remark 1.7.1 in [10]) provides the exact
rate of almost sure convergence of the number of upcrossings to the Brownian local time as follows:
Theorem 2.1. For every finite positive constant M > 0,
p
|U k (t, x) − ℓx (t)|
(2.1)
lim sup sup p
− 2 ℓ∗ (t) = 0 almost surely.
−k
k
k→∞ 0≤t≤M x∈R
2 log(2 )
A NOTE ON THE SHARP Lp -CONVERGENCE RATE OF UPCROSSINGS TO THE BROWNIAN LOCAL-TIME 3
where ℓ∗ (t) := supx∈R ℓx (t).
In this article, we are going to show a counterpart of Theorem 2.1 in the sense of Lp (P) for
1 ≤ p < ∞. More precisely, our main result reads as follows.
Theorem 2.2. Let T be a finite positive constant. For each η > 0, there exists a finite universal
positive constant C(η) such that
sup E sup sup
k≥1
0≤t≤T x∈R
|U k (t, x) − ℓx (t)|2+η
(2−k log(2k ))
2+η
2
1
η
≤ C(η)T 2 + 4 .
In the sequel, we denote Im := [−2m , 2m ] ⊂ R where m ≥ 1 is a given positive integer. We recall
the notion of q-variation of a real-valued function f : Im → R
kf kqIm ;q := sup
Π
X
|f (xi ) − f (xi−1 )|q ; 1 ≤ q < ∞,
xi ∈Π
where sup is taken over all partitions of the compact set Im (see e.g [7]). An important result due
to Feng and Zao [6] states that the Brownian local time has finite (2 + δ)-variation in the sense that
E sup0≤t≤T kℓ(t)k2+δ
Im ;2+δ < ∞ for every δ > 0. In this article, we show an upper bound for the number
upcrossings as follows. An almost immediate corollary of Theorem 2.2 is the following result:
Corollary 2.1. For any δ > 0 and m ≥ 1, there exists a finite positive constant C(δ, m) (which only
depends on δ and m) such that
1
δ
2+δ
2+4 .
sup E sup kU k (t)k2+δ
Im ;2+δ ≤ E sup kℓ(t)kIm ;2+δ + C(δ, m)T
k≥1
0≤t≤T
0≤t≤T
3. Proof of Theorem 2.2
It is not difficult to see that Theorem 2.1 will play a key role in the proof of Theorem 2.1. However,
the argument given by Koshnevisan in the proof of (2.1) is fully probabilistic in the sense that it
essentially relies on purely Borel-Cantelli’s-type arguments with none Lp estimates at hand. So we
need to adopt a rather different strategy. Thanks to the deep Burh¨older’s ideas [5] on moderate
functions, we shall construct an argument towards the proof of Theorem 2.2. The strategy is the
obtention of a so-called good-lambda inequality to get the desired Lp -rate of convergence. See Jacka [8]
and Bass [1] for further details.
The starting point of our analysis is the study of the scaling behavior of the following adapted
process
|U k (s, x) − ℓx (s)|2
; 0 ≤ t < ∞.
2−k log(2k )
0≤s≤t x∈R
The left-continuous version is given by
sup sup
|U k (s, x) − ℓx (s)|2
; 0 ≤ t < ∞.
2−k log(2k )
0≤s<t x∈R
J k (t) := sup sup
Since we are only interested on the bounded set [0, T ], we are going to stop J k as follows
F k (t) := J k (t ∧ T¯); 0 ≤ t < ∞
where T < T¯ < ∞.
Lemma 3.1. The following convergence holds
lim sup
sup
b→∞ k≥1 x∈R,λ>0
n
o
Px F k (λ2 ) > bλ = 0.
4
ALBERTO OHASHI AND ALEXANDRE B. SIMAS
Proof. Let us fix k ≥ 1. By the very definition,
sup
y∈R,λ>0
Py {F k (λ2 ) > bλ} ≤
≤
(3.1)
sup Py {F k (λ2 ) > bλ} + sup Py {F k (T¯) > bλ}
y∈R
0<λ2 <T¯
y∈R
λ2 ≥T¯
sup Py {F k (λ2 ) > bλ} + sup Py {F k (T¯) > bT¯ 1/2 }.
y∈R
0<λ2 <T¯
y∈R
By the very definition, if G is a Borel subset of Ω and y ∈ R then Py (G) = P(G − y) where G − y :=
{ω ∈ Ω; ω(·) + y ∈ G}. Now the Py law of F k (T¯) does not dependent on y ∈ R because of the
sup over all the initial conditions in R. Then, the almost sure convergence (2.1) and the fact that
ℓ∗ (T¯) < ∞ a.s yield
sup Py {F k (T¯) > bT¯ 1/2 } ≤ P{sup F r (T¯) > bT¯ 1/2 } → 0
y∈R
r≥1
as b → ∞ uniformly in k ≥ 1. It remains to estimate the first term in (3.1). We notice that for a
given λ > 0, the map s 7→ sλ is a bijection from [0, λ] onto [0, λ2 ] and hence
y
k
2
(3.2) sup sup P {F (λ ) > bλ}
=
0<λ2 <T¯ y∈R
sup sup P
y
0<λ2 <T¯ y∈R
(
sup sup
| √1λ U k (sλ,
0≤s<λ x∈R
√
√
√1 ℓ λx (sλ)|2
λ
2−k log(2k )
λx) −
)
>b .
Moreover, the Brownian motion scaling invariance yields
(3.3)
P
y
(
|U k (s, x) − ℓx (s)|2
>b
sup sup
2−k log(2k )
0≤s<λ x∈R
)
=P
y
(
sup sup
| √1λ U k (sλ,
0≤s<λ x∈R
√
√
√1 ℓ λx (sλ)|2
λ
2−k log(2k )
λx) −
Summing up identities (3.2) and (3.3), the fact that the Py law of sup0≤s<λ supx∈R
does not depend on y ∈ R and using (2.1), we do have
sup
y∈R
0<λ2 <T¯
y
k
2
P {F (λ ) > bλ}
≤
≤
(
|U k (s, x) − ℓx (s)|2
>b
P
sup sup
2−k log(2k )
0≤s≤T¯ 1/2 x∈R
(
)
>b .
|U k (s,x)−ℓx (s)|2
2−k log(2k )
)
|U r (s, x) − ℓx (s)|2
>b
P sup sup sup
2−r log(2r )
r≥1 0≤s≤T¯1/2 x∈R
)
→0
as b → ∞ uniformly in k ≥ 1. This concludes the proof.
In the sequel, let θt : Ω → Ω be the shift operator defined by θs ω := ω(· + s); ω ∈ Ω. For a given
adapted process {H(t); t ≥ 0}, we recall that H(·) ◦ θs (ω) := H(ω, · + s) for each ω ∈ Ω and s ≥ 0.
Lemma 3.2. For each k ≥ 1, the functional F k is an F-adapted non-decreasing functional with leftcontinuous paths which satisfies the following sub-additivity relation: For every 0 ≤ s ≤ t < ∞, we
have
F k (t) − F k (s) ≤ F k (t − s) ◦ θ(s) a.s
(3.4)
Proof. The fact that F k is adapted with left-continuous and non-decreasing paths is obvious. If
T¯ ≤ s < t, then (3.4) trivially holds. Now, if 0 ≤ s < t ≤ T¯, we clearly have
|U k (r, x) − ℓx (r)|2
|U k (r, x) − ℓx (r)|2
|U k (r, x) − ℓx (r)|2
≤ sup sup
+ sup sup
.
−k
k
−k
k
2 log(2 )
2 log(2 )
2−k log(2k )
0≤r<t x∈R
0≤r<s x∈R
s≤r<t x∈R
sup sup
A NOTE ON THE SHARP Lp -CONVERGENCE RATE OF UPCROSSINGS TO THE BROWNIAN LOCAL-TIME 5
Observe that the functional F k only depends on the time variable, and does not depend on the space
variable. Therefore, from the definition of the shift operator, we obtain:
|U k (r, x) − ℓx (r)|2
|U k (r, x) − ℓx (r)|2
= sup sup
◦ θs .
−k
k
2 log(2 )
2−k log(2k )
0≤r<t−s x∈R
s≤r<t x∈R
sup sup
This concludes the proof.
We are now able to prove Theorem 2.2. The idea is to find a good-lambda inequality (see e.g [8], [1])
for our functional F k . For this purpose, we fix k ≥ 1, η > 0 and for a given λ > 0, let us define
Jλ := inf{t ≥ 0; F k (t) > λ}. We also fix β > 1 and δ > 0. Since F k is left-continuous, then
F k (Jλ −) = F k (Jλ ) ≤ λ a.s. By using the sub-additive property of F k given in Lemma 3.2, the strong
Markov property of the Brownian motion, the non-decreasing and left-continuous paths of F k , we
shall find a good-lambda inequality as follows
P{F k (T¯) > βλ, (T¯ )1/2 ≤ δλ}
≤
P{F k (T¯ ) − F k (Jλ ) > (β − 1)λ, (T¯) ≤ δ 2 λ2 }
≤
P{F k (Jλ + δ 2 λ2 ) − F k (Jλ ) ≥ (β − 1)λ, Jλ < T¯}
≤
P{F k (δ 2 λ2 ) ◦ θJλ ≥ (β − 1)λ, Jλ < T¯}
=
Z
=
Z
≤
sup Px {F k (δ 2 λ2 ) ≥ (β − 1)λ}P{Jλ < T¯}
≤
n
β −1 o k ¯
sup Px F k (η 2 ) >
η P F (T ) > λ .
2δ
x∈R
{Jλ <T¯ }
{Jλ <T¯ }
P F k (δ 2 λ2 ) ◦ θJλ ≥ (β − 1)λ|FJλ dP
PB(Jλ ) F k (δ 2 λ2 ) ≥ (β − 1)λ dP
x∈R
η>0
Now, from Lemma 3.1, we shall take δ small enough in such way that
n
β−1 o
η
sup Px F k (η 2 ) >
2δ
x∈R
η>0
is small uniformly in k ≥ 1. Then, by applying Lemma 7.1 given in Burkh¨older [5] on the moderate
η
function x 7→ x1+ 2 (x ≥ 0), we get an universal constant which only depends on η such that
η
1
η
E|F k (T¯)|1+ 2 ≤ C(η)(T¯) 2 + 4 ∀k ≥ 1.
(3.5)
Since T¯ > T is arbitrary, then (3.5) concludes the proof of Theorem 2.2.
3.1. Proof of Corollary 2.1. Let us now give the proof of Corollary 2.1. In the sequel, we fix
δ > 0, m ≥ 1 and for a given partition Π = {xi }N
i=0 of the interval Im , let us define the following
subset Λ(Π, k) := {xi ∈ Π; (jk (xi ) − jk (xi−1 ))2−k > 0}. We notice that # Λ(Π, k) ≤ 22k+m for every
partition Π of Im . We readily see that
(3.6)
X
xi ∈Π
|U k (t, xi ) − U k (t, xi−1 )|2+δ ≤
X
xi ∈Λ(Π,k)
|U k (t, xi ) − U k (t, xi−1 )|2+δ ,
6
ALBERTO OHASHI AND ALEXANDRE B. SIMAS
for every partition Π of Im . By writing |U k (t, xi ) − U k (t, xi−1 )| = |U k (t, xi ) − ℓxi (t) + ℓxi (t) −
ℓxi−1 (t) + ℓxi−1 (t) − U k (t, xi−1 )|; xi ∈ Λ(Π, k) and applying the standard inequality |α − β|2+δ ≤
21+δ {|α|2+δ + |β|2+δ }; α, β ∈ R, we get from (3.6)
E sup kU k (t)k2+δ
Im ;2+δ
0≤t≤T
(3.7)
X
≤ CE sup kℓx (t)k2+δ
Im ;2+δ + CE sup sup
0≤t≤T
0≤t≤T
+ CE sup sup
0≤t≤T
Π
X
Π
|U k (t, xi ) − ℓxi (t)|2+δ
xi ∈Λ(Π,k)
|U k (t, xi−1 ) − ℓxi−1 (t)|2+δ ,
xi ∈Λ(Π,k)
for a constant C which only depends on δ > 0. An inspection in the proof of Lemma 2.1 in Feng and
Zao [6] yields E sup0≤t≤T kℓx (t)k2+δ
Im ;2+δ < ∞. Now,
X
sup sup
0≤t≤T
Π
|U k (t, xi ) − ℓxi (t)|2+δ ≤ 22k+m sup |U k (t, x) − ℓx (t)|2+δ a.s k ≥ 1
x∈Im
0≤t≤T
xi ∈Λ(Π,k)
and hence from Theorem 2.2, we have
E sup sup
0≤t≤T
Π
X
|U k (t, xi ) − ℓxi (t)|2+δ
1
δ
δ
≤
22k+m C(δ)T 2 + 4 (2−k klog(2))1+ 2
≤
2m+1 log(2)1+ 2 T 2 + 4 sup 2
xi ∈Λ(Π,k)
δ
1
δ
−rδ
2
δ
r1+ 2
r≥1
≤
1
δ
C(δ, m)T 2 + 4 < ∞,
for some constant c(δ, m) which only depends on m, δ. The other term in (3.7) can be treated similarly.
This concludes the proof of Corollary 2.1.
References
Lp -inequalities
[1] Bass, R. (1987).
for funcitonals of the Brownian motion. S´
eminaire de Probabilit´
es. Lecture Notes
in Math. 21, 206-217.
[2] Bass, R. and Koshnevisan, V. Strong approximations to Brownian local time. Sem. in Stoch. Proc. 1992, 43-65,
Birkhaser, (E. inlar, K. L. Chung, M. J. Sharp, R. F. Bass and K. Burdzy, Ed.’s).
[3] Blandine, B. and Vallois,P. (2008). Approximation via regularization of the local time of semimartingales and
Brownian motion. Stochastic Process. Appl, 118, 11, 2058-2070.
[4] Borodin, A.N (1986). On the character of convergence to Brownian local time.Z. Wahrch.Verw. Gebiete. 72, 231250.
[5] Burkholder, D. (1973). Distribution function inequalities for martingales. Ann.Probab., 1, 19-42.
[6] Feng, C. and Zhao, H. (2006). Two-parameter p, q-variation Paths and Integrations of Local-Times. Potential Anal,
25, 165-204
[7] Friz, P. Victoir, N. Multidimensional stochastic processes as rough paths. Theory and Applications. Cambridge
University Press. 2011.
[8] Jacka, S.D. (1989). A note on the good lambda inequalities. S´
eminaire de Probabilit´
es. Lecture Notes in Math. 23,
57-65.
[9] Karatzas, I. and Shreve, S. Brownian motion and stochastic calculus. Springer-Verlag.
[10] Khoshnevisan, D. (1994). Exact Rates o Convergence to Brownian Local Time. Ann. Probab.22, 3, 1295-1330.
[11] Knight, F. (1963). Random walk and a sojourn density process of Brownnian motion. Trans. Amer. Math. Soc,
109, 56-86.
[12] Knight, F. (1997). Approximation of stopped Brownian local time by diadic crossing chains. Stochastic Process.
Appl, 66, 2, 253-270.
[13] Le˜
ao, D. and Ohashi, A. (2013). Weak approximations for Wiener functionals. Ann. Appl. Probab, 23, 4, 1660-1691.
[14] Le˜
ao, D. Ohashi, A. and Simas, A. B. Weak Functional Itˆ
o Calculus and Applications. arXiv:1408.1423
[15] Ohashi, A. and Simas, A. B. A maximal inequaity for 2D Young integral based on bivariations. arXiv:1408.1428.
[16] Williams, D. (1977). Levy’s downcrossing theorem, Z. Wahrscheinlichkeitstheorie und Verw. Gebiete,40, 2, 157158.
A NOTE ON THE SHARP Lp -CONVERGENCE RATE OF UPCROSSINGS TO THE BROWNIAN LOCAL-TIME 7
[17] Young, L.C. (1937). General inequalities for Stieltjes integrals and the convergence of Fourier series. Mathematische
Annalen, 581-612.
´ tica, Universidade Federal da Para´ıba, 13560-970, Joa
˜ o Pessoa - Para´ıba, Brazil
Departamento de Matema
E-mail address: [email protected]; [email protected]
´ tica, Universidade Federal da Para´ıba, 13560-970, Joa
˜ o Pessoa - Para´ıba, Brazil
Departamento de Matema
E-mail address: [email protected]